Jump to content

Red panda: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎Diseases: changed
→‎Behaviour and ecology: corr spelling + int link
Line 138: Line 138:


===Social spacing===
===Social spacing===
Adult pandas are generally [[Solitary animal|solitary]] and [[Territory (animal)|territorial]]. Individuals mark their home range or territorial boundaries with urine, faeces and secretions from the [[anal gland|anal]] and surrounding glands. Scent-marking occurs more on the ground, and males mark more often and for longer periods than females.<ref name="roberts+gittleman" /> In China's [[Wolong National Nature Reserve]], the [[home range]] of a radio-collared female was {{cvt|0.94|km2}}, while that of a male was {{cvt|1.11|km2}}.<ref name="Reid1991">{{cite journal|last1=Reid|first1=D. G. |last2=Jinchu|first2=H. |last3=Yan|first3=H. |name-list-style=amp |year=1991 |title=Ecology of the Red Panda ''Ailurus fulgens'' in the Wolong Reserve, China |journal=Journal of Zoology |volume=225 |issue=3 |pages=347–364 |doi=10.1111/j.1469-7998.1991.tb03821.x |url=https://www.researchgate.net/publication/221959383}}</ref> A one-year-long monitoring study of ten red pandas in eastern Nepal showed that the four males had median home ranges of {{cvt|1.73|km2}} and the six females of {{cvt|0.94|km2}} within a forest cover of at least {{cvt|19.2|ha}}. The females travelled {{cvt|419–841|m}} per day and the males {{cvt|660–1473|m}}. In the mating season from January to March, adults travelled a mean of {{cvt|795|m}} and subadults a mean of {{cvt|861|m}}.<ref name=Bista_al2021b>{{cite journal |author1=Bista, D. |name-list-style=amp |author2=Baxter, G. S. |author3=Hudson, N. J. |author4=Lama, S. T. |author5=Weerman, J. |author6=Murray, P. J. |year=2021 |title=Movement and dispersal of a habitat specialist in human-dominated landscapes: a case study of the Red Panda |journal=Movement Ecology |volume=9 |issue=1 |page=62 |doi=10.1186/s40462-021-00297-z |pmid=34906253 |pmc=8670026 |doi-access=free}}</ref> They all had larger home ranges in areas with low forest cover and reduced their activity in areas that were disturbed by people, livestock and dogs.<ref name=Bista_al2021>{{cite journal |author1=Bista, D. |name-list-style=amp |author2=Baxter, G. S. |author3=Hudson, N. J. |author4=Lama, S. T. |author5=Murray, P. J. |year=2021 |title=Effect of disturbances and habitat fragmentation on an arboreal habitat specialist mammal using GPS telemetry: a case of the red panda |journal=Landscape Ecology |pages=1–15 |doi=10.1007/s10980-021-01357-w |pmid=34720409 |pmc=8542365 |doi-access=free}}</ref>
Adult pandas are generally [[Solitary animal|solitary]] and [[Territory (animal)|territorial]]. Individuals mark their home range or territorial boundaries with urine, [[feces]] and secretions from the [[anal gland|anal]] and surrounding glands. Scent-marking occurs more on the ground, and males mark more often and for longer periods than females.<ref name="roberts+gittleman" /> In China's [[Wolong National Nature Reserve]], the [[home range]] of a radio-collared female was {{cvt|0.94|km2}}, while that of a male was {{cvt|1.11|km2}}.<ref name="Reid1991">{{cite journal|last1=Reid|first1=D. G. |last2=Jinchu|first2=H. |last3=Yan|first3=H. |name-list-style=amp |year=1991 |title=Ecology of the Red Panda ''Ailurus fulgens'' in the Wolong Reserve, China |journal=Journal of Zoology |volume=225 |issue=3 |pages=347–364 |doi=10.1111/j.1469-7998.1991.tb03821.x |url=https://www.researchgate.net/publication/221959383}}</ref> A one-year-long monitoring study of ten red pandas in eastern Nepal showed that the four males had median home ranges of {{cvt|1.73|km2}} and the six females of {{cvt|0.94|km2}} within a forest cover of at least {{cvt|19.2|ha}}. The females travelled {{cvt|419–841|m}} per day and the males {{cvt|660–1473|m}}. In the mating season from January to March, adults travelled a mean of {{cvt|795|m}} and subadults a mean of {{cvt|861|m}}.<ref name=Bista_al2021b>{{cite journal |author1=Bista, D. |name-list-style=amp |author2=Baxter, G. S. |author3=Hudson, N. J. |author4=Lama, S. T. |author5=Weerman, J. |author6=Murray, P. J. |year=2021 |title=Movement and dispersal of a habitat specialist in human-dominated landscapes: a case study of the Red Panda |journal=Movement Ecology |volume=9 |issue=1 |page=62 |doi=10.1186/s40462-021-00297-z |pmid=34906253 |pmc=8670026 |doi-access=free}}</ref> They all had larger home ranges in areas with low forest cover and reduced their activity in areas that were disturbed by people, livestock and dogs.<ref name=Bista_al2021>{{cite journal |author1=Bista, D. |name-list-style=amp |author2=Baxter, G. S. |author3=Hudson, N. J. |author4=Lama, S. T. |author5=Murray, P. J. |year=2021 |title=Effect of disturbances and habitat fragmentation on an arboreal habitat specialist mammal using GPS telemetry: a case of the red panda |journal=Landscape Ecology |pages=1–15 |doi=10.1007/s10980-021-01357-w |pmid=34720409 |pmc=8542365 |doi-access=free}}</ref>


===Diet and feeding===
===Diet and feeding===
Line 163: Line 163:


=== Diseases===
=== Diseases===
[[Feces|Faecal]] samples of red panda collected in Nepal contained parasitic [[protozoa]], [[amoebozoa]]ns, [[roundworm]]s, [[trematodes]] and [[tapeworms]].<ref>{{cite journal |author1=Lama, S. T. |name-list-style=amp |author2=Lama, R. P. |author3=Regmi, G. R. |author4=Ghimire, T. R. |year=2015 |title=Prevalence of intestinal parasitic infections in free-ranging Red Panda ''Ailurus fulgens'' Cuvier, 1825 (Mammalia: Carnivora: Ailuridae) in Nepal |journal=Journal of Threatened Taxa |volume=7 |issue=8 |pages=7460–7464 |doi=10.11609/JoTT.o4208.7460-4 |doi-access=free}}</ref><ref>{{cite journal |author1=Bista, D. |name-list-style=amp |author2=Shrestha, S. |author3=Kunwar, A. J. |author4=Acharya, S. |author5=Jnawali, S. R. |author6=Acharya, K. P. |year=2017 |title=Status of gastrointestinal parasites in Red Panda of Nepal |journal=PeerJ |volume=5 |page=e3767 |doi=10.7717/peerj.3767 |pmid=28894643 |pmc=5591639 |doi-access=free}}</ref> Roundworms, tapeworms and [[coccidia]] were also found in red panda scat collected in Rara and Langtang National Parks.<ref>{{cite journal |author1=Sharma, H. P. |name-list-style=amp |author2=Achhami, B. |year=2021 |title=Gastro‐intestinal parasites of sympatric Red Panda and livestock in protected areas of Nepal |journal=Veterinary Medicine and Science |volume= |issue=|pmid=34599791 |s2cid=238250774 |doi=10.1002/vms3.651 |doi-access=free}}</ref> Fourteen red pandas at the [[Knoxville Zoo]] suffered from severe [[ringworm]], so the tails of two were [[Amputation|amputated]].<ref>{{cite journal |author1=Kearns, K. S. |name-list-style=amp |author2=Pollock, C. G. |author3=Ramsay, E. C. |year=1999 |title=Dermatophytosis in Red Pandas (''Ailurus fulgens fulgens''): a review of 14 cases |journal=Journal of Zoo and Wildlife Medicine |volume=30 |issue=4 |pages=561–563 |jstor=20095922|pmid=10749446 }}</ref> [[Chagas disease]] was reported as the cause of death of a red panda kept in a [[Kansas]] zoo.<ref>{{cite journal |author1=Huckins, G. L. |name-list-style=amp |author2=Eshar, D. |author3=Schwartz, D. |author4=Morton, M. |author5=Herrin, B. H. |author6=Cerezo, A. |author7=Yabsley, M. J. |author8=Schneider, S. M. |year=2019 |title=''Trypanosoma cruzi'' infection in a zoo-housed Red Panda in Kansas |journal=Journal of Veterinary Diagnostic Investigation |volume=31 |issue=5 |pages=752–755 |doi=10.1177/1040638719865926 |pmid=31342874 |pmc=6727118 |doi-access=free}}</ref> ''[[Amdoparvovirus]]'' was detected in the scat of six red pandas in the [[Sacramento Zoo]].<ref>{{cite journal |author1=Alex, C. E. |name-list-style=amp |author2=Kubiski, S. V. |author3=Li, L. |author4=Sadeghi, M. |author5=Wack, R. F. |author6=McCarthy, M. A. |author7=Pesavento, J. B. |author8=Delwart, E. |author9=Pesavento, P. A. |year=2018 |title=''Amdoparvovirus'' infection in Red Pandas (''Ailurus fulgens'') |journal=Veterinary Pathology |volume=55 |issue=4 |pages=552–561 |doi=10.1177/0300985818758470 |pmid=29433401 |doi-access=free}}</ref> Eight captive red pandas in a Chinese zoo suffered from [[shortness of breath]] and [[fever]] shortly before they died of [[pneumonia]]; [[autopsy]] revealed that they had antibodies to the protozoans ''[[Toxoplasma gondii]]'' and ''[[Sarcocystis]]'' species indicating that they were [[intermediate hosts]].<ref>{{cite journal |author1=Yang, Y. |name-list-style=amp |author2=Dong, H. |author3=Su, R. |author4=Li, T. |author5=Jiang, N. |author6=Su, C. |author7=Zhang, L. |year=2019 |title=Evidence of Red Panda as an intermediate host of ''Toxoplasma gondii'' and ''Sarcocystis'' species |journal=International Journal for Parasitology: Parasites and Wildlife |volume=8 |pages=188–191 |doi=10.1016/j.ijppaw.2019.02.006 |pmid=30891398 |pmc=6403407 |doi-access=free}}</ref> A captive red panda in the [[Chengdu Research Base of Giant Panda Breeding]] died of unknown reasons; an autopsy showed that its [[kidney]]s, [[liver]] and [[lung]]s were damaged by a bacterial infection caused by ''[[Escherichia coli]]''.<ref>{{cite journal |author1=Liu, S. |name-list-style=amp |author2=Li, Y. |author3=Yue, C. |author4=Zhang, D. |author5=Su, X. |author6=Yan, X. |author7=Yang, K. |author8=Chen, X. |author9=Zhuo, G. |author10=Cai, T. |author11=Liu, J. |author12=Peng, X. |author13=Huo, R. |year=2020 |title=Isolation and characterization of uropathogenic ''Escherichia coli'' (UPEC) from Red Panda (''Ailurus fulgens'') |journal=BMC Veterinary Research |volume=16 |issue=1 |page=404 |doi=10.1186/s12917-020-02624-9 |pmid=33109179 |pmc=7590469 |doi-access=free}}</ref>
Fecal samples of red panda collected in Nepal contained parasitic [[protozoa]], [[amoebozoa]]ns, [[roundworm]]s, [[trematodes]] and [[tapeworms]].<ref>{{cite journal |author1=Lama, S. T. |name-list-style=amp |author2=Lama, R. P. |author3=Regmi, G. R. |author4=Ghimire, T. R. |year=2015 |title=Prevalence of intestinal parasitic infections in free-ranging Red Panda ''Ailurus fulgens'' Cuvier, 1825 (Mammalia: Carnivora: Ailuridae) in Nepal |journal=Journal of Threatened Taxa |volume=7 |issue=8 |pages=7460–7464 |doi=10.11609/JoTT.o4208.7460-4 |doi-access=free}}</ref><ref>{{cite journal |author1=Bista, D. |name-list-style=amp |author2=Shrestha, S. |author3=Kunwar, A. J. |author4=Acharya, S. |author5=Jnawali, S. R. |author6=Acharya, K. P. |year=2017 |title=Status of gastrointestinal parasites in Red Panda of Nepal |journal=PeerJ |volume=5 |page=e3767 |doi=10.7717/peerj.3767 |pmid=28894643 |pmc=5591639 |doi-access=free}}</ref> Roundworms, tapeworms and [[coccidia]] were also found in red panda scat collected in Rara and Langtang National Parks.<ref>{{cite journal |author1=Sharma, H. P. |name-list-style=amp |author2=Achhami, B. |year=2021 |title=Gastro‐intestinal parasites of sympatric Red Panda and livestock in protected areas of Nepal |journal=Veterinary Medicine and Science |volume= |issue=|pmid=34599791 |s2cid=238250774 |doi=10.1002/vms3.651 |doi-access=free}}</ref> Fourteen red pandas at the [[Knoxville Zoo]] suffered from severe [[ringworm]], so the tails of two were [[Amputation|amputated]].<ref>{{cite journal |author1=Kearns, K. S. |name-list-style=amp |author2=Pollock, C. G. |author3=Ramsay, E. C. |year=1999 |title=Dermatophytosis in Red Pandas (''Ailurus fulgens fulgens''): a review of 14 cases |journal=Journal of Zoo and Wildlife Medicine |volume=30 |issue=4 |pages=561–563 |jstor=20095922|pmid=10749446 }}</ref> [[Chagas disease]] was reported as the cause of death of a red panda kept in a [[Kansas]] zoo.<ref>{{cite journal |author1=Huckins, G. L. |name-list-style=amp |author2=Eshar, D. |author3=Schwartz, D. |author4=Morton, M. |author5=Herrin, B. H. |author6=Cerezo, A. |author7=Yabsley, M. J. |author8=Schneider, S. M. |year=2019 |title=''Trypanosoma cruzi'' infection in a zoo-housed Red Panda in Kansas |journal=Journal of Veterinary Diagnostic Investigation |volume=31 |issue=5 |pages=752–755 |doi=10.1177/1040638719865926 |pmid=31342874 |pmc=6727118 |doi-access=free}}</ref> ''[[Amdoparvovirus]]'' was detected in the scat of six red pandas in the [[Sacramento Zoo]].<ref>{{cite journal |author1=Alex, C. E. |name-list-style=amp |author2=Kubiski, S. V. |author3=Li, L. |author4=Sadeghi, M. |author5=Wack, R. F. |author6=McCarthy, M. A. |author7=Pesavento, J. B. |author8=Delwart, E. |author9=Pesavento, P. A. |year=2018 |title=''Amdoparvovirus'' infection in Red Pandas (''Ailurus fulgens'') |journal=Veterinary Pathology |volume=55 |issue=4 |pages=552–561 |doi=10.1177/0300985818758470 |pmid=29433401 |doi-access=free}}</ref> Eight captive red pandas in a Chinese zoo suffered from [[shortness of breath]] and [[fever]] shortly before they died of [[pneumonia]]; [[autopsy]] revealed that they had antibodies to the protozoans ''[[Toxoplasma gondii]]'' and ''[[Sarcocystis]]'' species indicating that they were [[intermediate hosts]].<ref>{{cite journal |author1=Yang, Y. |name-list-style=amp |author2=Dong, H. |author3=Su, R. |author4=Li, T. |author5=Jiang, N. |author6=Su, C. |author7=Zhang, L. |year=2019 |title=Evidence of Red Panda as an intermediate host of ''Toxoplasma gondii'' and ''Sarcocystis'' species |journal=International Journal for Parasitology: Parasites and Wildlife |volume=8 |pages=188–191 |doi=10.1016/j.ijppaw.2019.02.006 |pmid=30891398 |pmc=6403407 |doi-access=free}}</ref> A captive red panda in the [[Chengdu Research Base of Giant Panda Breeding]] died of unknown reasons; an autopsy showed that its [[kidney]]s, [[liver]] and [[lung]]s were damaged by a bacterial infection caused by ''[[Escherichia coli]]''.<ref>{{cite journal |author1=Liu, S. |name-list-style=amp |author2=Li, Y. |author3=Yue, C. |author4=Zhang, D. |author5=Su, X. |author6=Yan, X. |author7=Yang, K. |author8=Chen, X. |author9=Zhuo, G. |author10=Cai, T. |author11=Liu, J. |author12=Peng, X. |author13=Huo, R. |year=2020 |title=Isolation and characterization of uropathogenic ''Escherichia coli'' (UPEC) from Red Panda (''Ailurus fulgens'') |journal=BMC Veterinary Research |volume=16 |issue=1 |page=404 |doi=10.1186/s12917-020-02624-9 |pmid=33109179 |pmc=7590469 |doi-access=free}}</ref>


== Threats ==
== Threats ==

Revision as of 09:35, 28 February 2022

Red panda
CITES Appendix I (CITES)[1]
Scientific classification Edit this classification
Domain: Eukaryota
Kingdom: Animalia
Phylum: Chordata
Class: Mammalia
Order: Carnivora
Family: Ailuridae
Genus: Ailurus
F. Cuvier, 1825
Species:
A. fulgens
Binomial name
Ailurus fulgens
F. Cuvier, 1825
Subspecies

A. f. fulgens F. Cuvier, 1825
A. f. styani Thomas, 1902[2]

Map showing the range of the red panda in red
Range of the red panda

The red panda (Ailurus fulgens), also known as the lesser panda, is a small mammal native to the eastern Himalayas and southwestern China. It has dense reddish-brown fur, white-lined ears, a mostly white muzzle and a ringed tail. Its head-to-body length is 51–63.5 cm (20.1–25.0 in) with a 28–48.5 cm (11.0–19.1 in) tail, and it weighs between 3.2 and 15 kg (7.1 and 33.1 lb). It is well adapted to climbing due to its flexible joints and curved semi-retractile claws.

The red panda was first described in 1825. The two currently recognised subspecies, the Himalayan and the Chinese red panda, genetically diverged about 250,000 years ago. The red panda's place on the evolutionary tree has been debated, but modern genetic evidence places it in close affinity with raccoons and weasels. It is not closely related with the giant panda (Ailuropoda melanoleuca), which is a bear, though both possess elongated wrist bones or "false thumbs" used for grasping bamboo. The evolutionary lineage of the red panda stretches back around 25 to 18 million years ago, as indicated by extinct fossil relatives found in Eurasia and North America.

The red panda inhabits coniferous forests as well as temperate broadleaf and mixed forests, favouring steep slopes with dense bamboo cover close to water sources. It is solitary and largely arboreal. It feeds mainly on bamboo shoots and leaves, but also on fruits and blossoms. Red pandas mate in early spring, with the females giving birth to litters of up to four cubs in summer. It is threatened by poaching as well as destruction and fragmentation of habitat due to deforestation. The species has been listed as Endangered on the IUCN Red List since 2015. It is protected in all range countries.

Community-based conservation programmes have been initiated in Nepal, Bhutan and northeastern India; in China, it benefits from nature conservation projects. The International Red Panda Day is celebrated annually in September. Regional captive breeding programmes for the red panda have been established in zoos around the world. It is featured in animated movies, video games, comic books and as the namesake of companies and music bands.

Etymology

The name "panda" is thought to have originated from the red panda's local Nepali name पञ्जा pajā "claw" or पौँजा paũjā "paw".[3][4] In English, it was simply called "panda"; when the giant panda (Ailuropoda melanoleuca) was formally described and named in 1869, it became known as the "red panda" or "lesser panda" to distinguish it from the larger animal.[4] The genus name Ailurus is adopted from the ancient Greek word αἴλουρος (ailouros), meaning "cat".[5] The specific epithet fulgens is Latin for "shining, bright".[6][4]

Taxonomy

Illustration of the red panda by Joseph Smit, 1869

The red panda was classified and formally described in 1825 by Frederic Cuvier, who gave it its current scientific name Ailurus fulgens. Cuvier's description was based on zoological specimens, including skin, paws, jawbones and teeth "from the mountains north of India", as well as an account by Alfred Duvaucel.[7][8] The red panda was described earlier by Thomas Hardwicke in 1821, but his paper was published only six years later.[4]

In 1847, Brian Houghton Hodgson described a red panda from the Himalayas, for which he proposed the name Ailurus ochraceus.[9] For a long time, Hodgson's account was the only information available about the red panda's behaviour in the wild.[4] In 1902, Oldfield Thomas described a skull of a male red panda specimen collected in Sichuan by Frederick William Styan under the name Ailurus fulgens styani.[2]

Subspecies and species

The modern red panda is the only recognised species of the genus Ailurus. It is traditionally divided into two subspecies: the Himalayan red panda (A. f. fulgens) and the Chinese red panda (A. f. styani). The Himalayan subspecies has a straighter profile, a lighter coloured forehead and ochre-tipped hairs on the lower back and rump. The Chinese subspecies has a more curved forehead, steeper muzzle slope, a darker coat with a redder, less white face and more contrast between the tail rings.[10]

In 2020, results of a genetic analysis of red panda samples showed that the red panda populations in the Himalayas and China were separated about 250,000 years ago. The researchers suggested that the two subspecies should be treated as distinct species. Red pandas in southeastern Tibet and northern Myanmar were found to be part of styani, while those of southern Tibet were of fulgens (sensu stricto).[11] DNA sequencing of 132 red panda faecal samples collected in Northeast India and China also showed two distinct clusters indicating that the Siang River in Arunachal Pradesh constitutes the boundary between the Himalayan and Chinese red pandas.[12] They probably diverged due to glaciation events on the southern Tibetan Plateau in the Pleistocene.[13]

Phylogeny

The placement of the red panda on the evolutionary tree has been debated. In the first half of the 20th century, various scientists placed it in the family Procyonidae with raccoons and their allies. At the time, most prominent biologists also considered the red panda to be related to the giant panda and classified both in the subfamily Ailurinae within Procyonidae. The giant panda would eventually be found to be a bear. A 1982 study examined the similarities and differences in the skull between the red panda and the giant panda, other bears and procyonids, and placed the species in its own family Ailuridae. The author of the study considered the red panda to be more closely related to bears.[10]

A 1995 mitochondrial DNA analysis revealed that the red panda has close affinities with procyonids.[14] Further genetic studies have placed the red panda within the clade Musteloidea, which also includes Procyonidae, Mustelidae (weasels and relatives) and Mephitidae (skunks and relatives). The following cladogram is based on the molecular phylogeny of six genes,[15] with the musteloids updated following a multigene analysis.[16]

Caniformia

Canidae (dogs and other canines) African golden wolf

Arctoidea

Ursidae (bears) American black bear

Pinnipedia (seals) Common seal

Musteloidea

Mephitidae (skunks) Striped skunk

Ailuridae (red panda) Red panda

Procyonidae (raccoons and allies) Common raccoon

Mustelidae (weasels and allies) European polecat

Fossil record

Reconstructed skull and head of Simocyon

The family Ailuridae appears to have originated in Europe sometime during the Late Oligocene or Early Miocene, about 25 to 18 million years ago. The earliest member Amphictis is known from its 10 cm (4 in) skull and may have been around the same size as the modern species. Its dentition consists of pointed premolars, relatively sharp-edged carnassials (P4 and m1) and molars with grinding surfaces (M1, M2 and m2), suggesting that it had a generalized carnivorous diet. Its placement within Ailuridae is based on the lateral grooves on its canine teeth. Other early or basal aliruds include Alopecocyon and Simocyon, whose fossils have been found throughout Eurasia and North America dating from the Middle Miocene, the latter of which survived into the Early Pliocene. Both have similar teeth to Amphictis and thus had a similar diet.[17] The puma-sized Simocyon was likely a tree-climber and shares a "false thumb"—an extended wrist bone—with the modern species, suggesting the appendage was an adaptation to arboreal locomotion and not to feed on bamboo.[17][18]

Later and more advanced ailruds are classified in the subfamily Ailurinae and are known as the "true" red pandas. These animals were smaller and more adapted for an omnivorous or herbivorous diet. The earliest known true panda is the species Magerictis imperialensis from the Middle Miocene of Spain and known only from a single tooth, a lower second molar. The tooth shows both ancestral and new characteristics having a relatively low and uncomplex crown but also an elongated crushing surface and well-differentiated tooth cusps like later species.[19] Later ailurines include Pristinailurus bristoli of Late Miocene-Early Pliocene eastern North America[19][20] and species of the genus Parailurus which first appear in Early Pliocene Europe, spreading across Eurasia into North America.[19][21] These animals are likely to be part of a sister taxon to the lineage of the modern panda. In contrast to the herbivorous modern species, these ancient pandas were likely omnivores, possessing many cusps on the molars but retaining sharp premolars.[19][22][20]

The earliest fossil record of the modern genus Ailurus date no earlier than the Pleistocene and appears to have been limited to Asia. The modern panda's lineage became adapted for a specialized bamboo diet, having molar-like premolars and more highly crowned cusps.[19] The false thumb would secondarily gain a function in feeding.[17][18]

Genomics

Analysis of 53 red panda samples from Sichuan and Yunnan showed a high level of genetic diversity.[23] The full genome of the red panda was sequenced in 2017. Researchers have compared it to the genome of the giant panda to learn the genetics of convergent evolution, as both species have false thumbs and are adapted for a specialized bamboo diet despite having the digestive system of a carnivore. Both pandas show modifications to certain limb development genes (DYNC2H1 and PCNT), which may play roles in the development of the thumbs.[24] In switching from a carnivorous to a herbivorous diet, both species have reactivated taste receptor genes used for detecting bitterness, though the specific genes are different.[25]

Characteristics

Red panda skull
Red panda face

The red panda's coat is mainly red or orange-brown with a black belly and legs. The face is mostly white and has red marks that stretch from the side angle of the eyes to the corners of the mouth. The inside of the ears are covered in white fur with a red patch in the centre.[26] Its bushy tail has alternating rings of red and buff.[27][26] The colouration appears to serve as camouflage in a habitat with red moss- and white lichen-covered trees. The fur consists of coarse guard hairs with a soft dense, woolly undercoat.[27] The guard hairs on the back have a circular cross-section and are 47–56 mm (1.9–2.2 in) long. It has moderately long whiskers around the mouth, lower jaw and chin.[26]

The red panda has a head-body length of 51–63.5 cm (20.1–25.0 in) with a 28–48.5 cm (11.0–19.1 in) tail. The Himalayan red panda is recorded to weigh 3.2–9.4 kg (7.1–20.7 lb), while the Chinese red panda weighs 4–15 kg (8.8–33.1 lb) for females and 4.2–13.4 kg (9.3–29.5 lb) for males.[26] The panda has a relatively small head with a reduced snout and triangular ears, though proportionally larger than in similarly sized raccoons, while the limbs are nearly equal in length.[26][27] The red panda has five curved digits on each foot, which end in curved semi-retractile claws that aid in climbing.[27] The pelvis and hindlimbs have flexible joints, adaptations for an arboreal quadrupedal lifestyle.[28] While not prehensile, the tail acts as support and counterbalance when climbing.[27]

The forepaws possess a "false thumb", which is an extension of a wrist bone, the radial sesamoid found in many carnivorans. This thumb allows the animal to hold onto bamboo stalks and separate leaves, and both the digits and wrist bones give the red panda remarkable dexterity. The red panda shares this feature with the giant panda, which has a larger sesamoid that is more compressed at the sides. In addition, the red panda's sesamoid has a more concave tip while the giant panda's hooks in the middle.[29]

Its skull is wide, and its lower jaw is robust.[27][26] However, because it eats the less fibrous parts of bamboo, the leaves and stems, it has less-developed chewing muscles than the giant panda. The digestive tract of the red panda is also typical of a carnivore, being fairly short, at only 4.2 times its body length, with a simple stomach, no clear distinction between the ileum and the colon, and no caecum.[26] Microbes in its gut may play a role in its processing of bamboo; the microbiota community in the red panda is less diverse than in other mammals.[30]

Distribution and habitat

Red panda in Neora Valley National Park

The red panda is distributed from western Nepal, the states of Sikkim, West Bengal and Arunachal Pradesh in India, Bhutan and southern Tibet to northern Myanmar and China's Sichuan and Yunnan provinces.[1] The global potential habitat of the red panda has been estimated to comprise 47,100 km2 (18,200 sq mi) at most; this habitat is located in the temperate climate zone of the Himalayas with a mean annual temperature range of 18–24 °C (64–75 °F).[31] Throughout this range, it has been recorded at elevations of 2,000–4,300 m (6,600–14,100 ft).[32][33][34][35][36]

Habitat of the red panda
Country Estimated size[31]
Nepal 22,400 km2 (8,600 sq mi)
China 13,100 km2 (5,100 sq mi)
India 5,700 km2 (2,200 sq mi)
Myanmar 5,000 km2 (1,900 sq mi)
Bhutan 900 km2 (350 sq mi)
Total 47,100 km2 (18,200 sq mi)

In Nepal, it lives in six protected area complexes within the Eastern Himalayan broadleaf forests ecoregion.[34] The westernmost records to date were obtained in three community forests in Kalikot District in 2019.[37] Panchthar and Ilam Districts represent its easternmost range in the country, where its habitat in forest patches is surrounded by villages, livestock pastures and roads.[38] The metapopulation in protected areas and wildlife corridors in the Kangchenjunga landscape of Sikkim and northern West Bengal is partly connected through old-growth forests outside protected areas.[39] Forests in this landscape are dominated by Himalayan oaks (Quercus lamellosa and Q. semecarpifolia), Himalayan birch (Betula utilis), Himalayan fir (Abies densa), Himalayan maple (Acer caesium) with bamboo, Rhododendron and some black juniper (Juniperus indica) shrub growing in the understoreys.[32][40][41][42] Records in Bhutan, Arunachal Pradesh's Pangchen Valley, West Kameng and Shi Yomi districts indicate that it frequents habitats with Yushania and Thamnocalamus bamboo, medium-sized Rhododendron, Sorbus and Castanopsis trees.[33][43][44] In China, it inhabits the Hengduan Mountains subalpine conifer forests and Qionglai-Minshan conifer forests in the Hengduan, Qionglai, Xiaoxiang, Daxiangling and Liangshan Mountains in Sichuan.[45] In the adjacent Yunnan province, it was recorded only in the northwestern montane part.[46][47]

The red panda prefers microhabitats within 70–240 m (230–790 ft) of water sources.[48][49][50][51] Fallen logs and tree stumps are important habitat features, as they facilitate access to bamboo leaves.[52] Red pandas have been recorded to use steep slopes of more than 20° and stumps exceeding a diameter of 30 cm (12 in).[48][53] Red pandas observed in Phrumsengla National Park used foremost easterly and southerly slopes with a mean slope of 34° and a canopy cover of 66 percent that were overgrown with bamboo about 23 m (75 ft) in height.[49] In Dafengding Nature Reserve, it prefers steep south-facing slopes in winter and inhabits forests with bamboo 1.5–2.5 m (4 ft 11 in – 8 ft 2 in) tall.[54] In Gaoligongshan National Nature Reserve, it inhabits mixed coniferous forest with a dense canopy cover of more than 75 percent, steep slopes and a density of at least 70 bamboo plants/m2 (6.5 bamboo plants/sq ft).[55] In China's Fengtongzhai and Yele Natural Reserves, the red panda selects steep slopes and a high density of bamboo culms, fallen logs and stumps, whereas the giant panda prefers gentle slopes with taller bamboo but lower densities of culms, logs and stumps. Such niche separation lessens competition between the two bamboo-eating species.[48][52]

Behaviour and ecology

A red panda lies sleeping on a high branch of a tree, with tail stretched out behind and legs dangling on each side of the branch
Red panda sleeping on a tree

The red panda is difficult to observe in the wild.[56] Most of the knowledge about its behaviour was found out in studies carried out in zoos.[57] The red panda appears to be both nocturnal and crepuscular, sleeping in between periods of activity at night. It typically rests or sleeps in trees or other elevated spaces, stretched out on a branch with legs dangling when it is hot, and curled up with its tail over the face when it is cold. It is adapted for climbing and descends to the ground head-first with the hindfeet holding on to the middle of the tree trunk. It moves quickly on the ground by trotting or bounding. Its lifespan in captivity reaches 14 years.[27]

Social spacing

Adult pandas are generally solitary and territorial. Individuals mark their home range or territorial boundaries with urine, feces and secretions from the anal and surrounding glands. Scent-marking occurs more on the ground, and males mark more often and for longer periods than females.[27] In China's Wolong National Nature Reserve, the home range of a radio-collared female was 0.94 km2 (0.36 sq mi), while that of a male was 1.11 km2 (0.43 sq mi).[58] A one-year-long monitoring study of ten red pandas in eastern Nepal showed that the four males had median home ranges of 1.73 km2 (0.67 sq mi) and the six females of 0.94 km2 (0.36 sq mi) within a forest cover of at least 19.2 ha (47 acres). The females travelled 419–841 m (1,375–2,759 ft) per day and the males 660–1,473 m (2,165–4,833 ft). In the mating season from January to March, adults travelled a mean of 795 m (2,608 ft) and subadults a mean of 861 m (2,825 ft).[38] They all had larger home ranges in areas with low forest cover and reduced their activity in areas that were disturbed by people, livestock and dogs.[59]

Diet and feeding

Red panda feeding

The red panda is largely herbivorous and feeds primarily on bamboo, mainly the genera Phyllostachys, Sinarundinaria, Thamnocalamus and Chimonobambusa.[60] It also feeds on fruits, blossoms, acorns, eggs, birds and small mammals. It mainly eats the leaves of bamboo, which are often the only available food item in the winter and the most common food for the rest of the year.[61] In Wolong National Nature Reserve, leaves of Bashania fangiana were found in nearly 94 percent of analysed droppings, and its shoots were found in 59 percent of the droppings found in June.[58]

The diet of red pandas monitored at three sites in Singalila National Park for two years consisted of 40–83 percent Yushania maling and 51–91.2 percent Thamnocalamus spathiflorus bamboos[a] supplemented by bamboo shoots, Actinidia strigosa fruits and seasonal berries.[64] In this national park, red panda droppings also contained remains of silky rose and bramble fruit species in the summer season, Actinidia callosa in the post-monsoon season, and Merrilliopanax alpinus, whitebeam (Sorbus cuspidata) and tree rhododendron in both seasons. Droppings were found on 23 plant species including the stone oak (Lithocarpus pachyphyllus), Campbell's magnolia (Magnolia campbellii), chinquapin (Castanopsis tribuloides), Himalayan birch, Litsea sericea and the holly species Ilex fragilis.[65] In Nepal's Rara National Park, Thamnocalamus was found in 100 percent of droppings sampled, both before and after the monsoon.[66] Its summer diet in Dhorpatan Hunting Reserve also includes some lichens and barberries.[40] In Bhutan's Jigme Dorji National Park, red panda feces found in the fruiting season contained seeds of Himalayan ivy (Hedera nepalensis).[51]

The red panda grabs food with one of its front paws and usually eats sitting down or standing, but sometimes lays on its back. When foraging for bamboo, it grabs the plant by the culm and bends it down so the leaves are within reach of the jaws. It inserts them into the side and shears and chews them. It nips small food like blossoms, berries and small leaves with the incisors.[27] The red panda is a poor digester of bamboo, which passes through its gut in two to four hours. It hence selects the more nutritious plant matter, such as tender leaves and shoots, and consumes them in large quantities. It eats over 1.5 kg (3 lb 5 oz) of fresh leaves or 4 kg (9 lb) of fresh shoots in a day and can digest crude proteins and fats more easily than fibres and lignin in the bamboo leaves. Bamboo is most digestible in summer and fall but least in winter, and shoots are better digestible than leaves.[67]

Communication

Sounds of red panda twittering

At least seven different vocalisations have been recorded in the red panda, comprising growls, barks, squeals, hoots, bleats, grunts and twitters. Growling, barking, grunting and squealing are produced during fights and aggressive chasing. Hooting is made in response to being approached by another individual. Bleating is recorded after scent-marking and sniffing and males may bleat during courtship, particularly before mounting, while twittering is made by mating females.[68]

During both play fighting and aggressive fighting, individuals arch their backs and tails while slowly moving their heads up and down. They then turn their heads while jaw-clapping, move their heads side to side and raise a forepaw with an intent to strike. They stand on their hind legs and raise the forelimbs above the head before lunging. Two individuals "stare" each other from a distance.[27] Receptive females make tail-flicks and position themselves in a lordosis pose, with the front lowered and the back arched.[69]

Reproduction and parenting

Red panda tending its cub

Red pandas are "long-day" breeders, meaning that breeding occurs as the length of daylight increases following the winter solstice. Mating thus occurs mostly between January and March, with births taking place from May to August. For captive pandas in the southern hemisphere, reproduction is delayed by six months. Oestrous lasts a day, and females can enter oestrous multiple times a season, but the length of intervals between each cycle is not clear.[69]

As the breeding season begins, there are increased interactions between males and females, who will rest, move and feed close to each other. Oestrous females are observed to mark more often and more vigorously and males will sniff their anogenital region. Copulation involves the male mounting the female from behind and on top, though face-to-face matings as well as belly-to-back matings while lying on the sides have been observed. The male usually does not bite the female's neck but will grab her sides with his front paws. Mountings are 2–25 minutes long, and the couple may groom each other between mounting bouts.[69]

Gestation lasts about 158 days. Prior to giving birth, the female selects a denning site, such as a tree, log or stump hollow or rock crevice, and builds a nest using material from nearby.[56] Litters typically consist of one to four cubs that are born fully furred but blind. They are entirely dependent on their mother for the first three to four months until they emerge from the nest. They nurse for their first five months.[70] Mother and offspring stay together until the next breeding. Cubs reach their adult size at around 12 months and sexual maturity at around 18 months.[27] Two radio-collared cubs in eastern Nepal separated from their mothers at the age of 7–8 months and left their birth areas three weeks later. They reached new home ranges within 26–42 days and became residents after exploring them for 42–44 days.[38]

Diseases

Fecal samples of red panda collected in Nepal contained parasitic protozoa, amoebozoans, roundworms, trematodes and tapeworms.[71][72] Roundworms, tapeworms and coccidia were also found in red panda scat collected in Rara and Langtang National Parks.[73] Fourteen red pandas at the Knoxville Zoo suffered from severe ringworm, so the tails of two were amputated.[74] Chagas disease was reported as the cause of death of a red panda kept in a Kansas zoo.[75] Amdoparvovirus was detected in the scat of six red pandas in the Sacramento Zoo.[76] Eight captive red pandas in a Chinese zoo suffered from shortness of breath and fever shortly before they died of pneumonia; autopsy revealed that they had antibodies to the protozoans Toxoplasma gondii and Sarcocystis species indicating that they were intermediate hosts.[77] A captive red panda in the Chengdu Research Base of Giant Panda Breeding died of unknown reasons; an autopsy showed that its kidneys, liver and lungs were damaged by a bacterial infection caused by Escherichia coli.[78]

Threats

The primary threats to the red panda are destruction and fragmentation of habitat caused by multiple circumstances such as increasing human population, deforestation, illegal collection of non-timber forest products and poaching, disturbances by herders and livestock, lack of law enforcement and funding.[1] Small groups of animals with little opportunity for exchange between them face the risk of inbreeding, decreased genetic diversity, and even extinction. In addition, clearcutting for firewood or agriculture and hillside terracing removes old trees that provide maternal dens and decreases the ability of some bamboo species to regenerate.[79] The cut lumber stock in Sichuan alone reached 2,661,000 m3 (94,000,000 cu ft) in 1958–1960, and 3,597.9 km2 (1,389.2 sq mi) of red panda habitat were logged between the mid 1970s and late 1990s.[46]

Deforestation inhibits the dispersal of red pandas and leads to severe fragmentation of the population; trampling by livestock depresses bamboo growth.[80] Throughout Nepal, the red panda habitat outside protected areas is negatively affected by solid waste, livestock trails and herding stations, and people collecting firewood and medicinal plants.[40][81] Threats identified in Nepal's Lamjung District include grazing by livestock during seasonal transhumance, human-made forest fires and the collection of bamboo as cattle fodder in winter.[82] Vehicular traffic is a significant barrier to red panda movement between habitat patches.[59]

In Nepal's Taplejung District, red panda claws are used for treating epilepsy; its skin is used in rituals for treating sick people, making hats, scarecrows and decorating houses. Between 2008 and 2018, 121 skins were confiscated in the country.[83] In Myanmar, the red panda is threatened by hunting using guns and traps; since roads to the border with China were built starting in the early 2000s, red panda skins and live animals are traded and smuggled across the border.[36] In southwestern China, the red panda is hunted for its fur, especially for the highly valued bushy tails, from which hats are produced. The fur is used for local cultural ceremonies. At weddings, the bridegroom traditionally carries the hide. The "good-luck charm" red panda-tail hats are also used by local newlyweds. A 40 percent decrease in red panda populations has been reported in China over the last 50 years, and populations in western Himalayan areas are considered to be smaller.[46] Between 2005 and 2017, 35 live and seven dead red pandas were confiscated in Sichuan, and several traders were sentenced to 3–12 years of imprisonment. A month-long survey of 65 shops in nine Chinese counties in the spring of 2017 revealed only one in Yunnan offered hats made of red panda skins, and red panda tails were offered in an online forum.[84]

Conservation

The red panda is listed in CITES Appendix I and protected in all range countries; hunting is illegal. It has been listed as Endangered on the IUCN Red List since 2008 because the global population is estimated at 10,000 individuals, with a decreasing population trend. A large extent of its habitat is part of protected areas.[1]

Protected areas in red panda range countries
Country Protected areas
Nepal Api Nampa Conservation Area, Khaptad National Park, Rara National Park, Annapurna Conservation Area, Manaslu Conservation Area, Langtang National Park, Gaurishankar Conservation Area, Sagarmatha National Park, Makalu Barun National Park, Kanchenjunga Conservation Area[34]
India Khangchendzonga National Park, Singalila National Park, Varsey Rhododendron Sanctuary, Shingba Rhododendron Sanctuary, Fambong Lho Wildlife Sanctuary, Kyongnosla Alpine Sanctuary, Pangolakha Wildlife Sanctuary, Maenam Wildlife Sanctuary,[39] Namdapha National Park[85]
Bhutan Jigme Khesar Strict Nature Reserve, Jigme Dorji National Park, Wangchuck Centennial National Park, Jigme Singye Wangchuck National Park, Bumdeling Wildlife Sanctuary, Sakteng Wildlife Sanctuary, Phrumsengla National Park, Jomotsangkha Wildlife Sanctuary[33]
Myanmar Hkakaborazi National Park, Hponkanrazi Wildlife Sanctuary,[86] Imawbum National Park[36]
China Yarlung Tsangpo Grand Canyon Nature Reserve[87] and six more nature reserves in Tibet, eight in Yunnan and 32 in Sichuan[88]
Closeup look of red panda

A red panda anti-poaching unit and community-based monitoring have been established in Langtang National Park. Members of Community Forest User Groups also protect and monitor red panda habitats in other parts of Nepal.[89] Community outreach programs have been initiated in eastern Nepal using information boards, radio broadcasting and the annual International Red Panda Day in September; several schools endorsed a red panda conservation manual as part of their curricula.[90]

Since 2010, community-based conservation programmes have been initiated in 10 districts in Nepal that aim to help villagers reduce their dependence on natural resources through improved herding and food processing practices and alternative income possibilities. The Nepal government ratified a five-year Red Panda Conservation Action Plan in 2019.[91] From 2016 to 2019, 35 ha (86 acres) of high-elevation rangeland in Merak, Bhutan, was restored and fenced in cooperation with 120 herder families to protect the red panda forest habitat and improve communal pasture.[92] Villagers in Arunachal Pradesh established two community conservation areas to protect the red panda habitat from disturbance and exploitation of forest resources.[43] China has initiated several projects to protect its environment and wildlife, including Grain for Green, The Natural Forest Protection Project and the National Wildlife/Natural Reserve Construction Project. For the last project, the red panda is not listed as a key animal for protection but may benefit from the protection of the giant panda and golden snub-nosed monkey (Rhinopithecus roxellana), with which it overlaps in range.[88]

In captivity

Red panda at Symbio Wildlife Park

The London Zoo acquired two red pandas in 1869 and 1876 that were caught in Darjeeling. The Calcutta Zoo received a live red panda in 1877, the Philadelphia Zoo in 1906, and Artis and Cologne Zoos in 1908. In 1908, the first red panda cubs were born in an Indian zoo. In 1940, the San Diego Zoo imported four red pandas via India that had been caught in Nepal; their first litter was born in 1941. Cubs that were born later were sent to other zoos so that about 250 red pandas had been exhibited in zoos by 1969.[93] The Taronga Conservation Society started keeping red pandas in 1977.[94]

In 1978, the International Red Panda Studbook was set up, followed by the Red Panda European Endangered Species Programme in 1985. Members of international zoos ratified a global master plan for the captive breeding of the red panda in 1993. By the end of 2019, 182 European zoos kept 407 red pandas.[95] By late 2015, 219 red pandas lived in 42 zoos in Japan.[96] The Padmaja Naidu Himalayan Zoological Park participates in the Red Panda Species Survival Plan and kept about 25 red pandas by 2016.[97] Regional captive breeding programmes have also been established in North American, Australasian and South African zoos.[4]

Cultural significance

The red panda is depicted in a hunting scene of a Chinese Chou Dynasty scroll dating to the 13th century. In western Nepal, shamans of an ethnic group use their skin and fur in their ritual dresses and believe that it protects against evil spirits. Tribal people in Arunachal Pradesh and Yi people also believe that it brings good luck to wear red panda tails or hats made of its fur. People in central Bhutan consider red pandas to be reincarnations of Buddhist monks.[98]

A watercolour painting by an Indian artist dating to 1820 is among the earliest known paintings of the red panda.[99] The red panda was recognized as the state animal of Sikkim in the early 1990s and was the mascot of the Darjeeling Tea Festival.[79] Anthropomorphic red pandas feature in animated movies and TV series such as Bamboo Bears, Barbie as the Island Princess, the Kung Fu Panda franchise, Aggretsuko and Turning Red, and in several video games and comic books. The red panda is the namesake of the Firefox browser, and it has been used as the namesake of companies and music bands.[98]

Notes

  1. ^ Labelled Arundinaria maling and A. aristata respectively, which are junior synonyms of the species listed here.[62][63]

References

  1. ^ a b c d e Glatston, A.; Wei, F.; Than Zaw & Sherpa, A. (2017) [errata version of 2015 assessment]. "Ailurus fulgens". IUCN Red List of Threatened Species. 2015: e.T714A110023718. Retrieved 15 January 2022.
  2. ^ a b Thomas, O. (1902). "On the Panda of Sze-chuen". Annals and Magazine of Natural History. 7. X (57): 251–252. doi:10.1080/00222930208678667.
  3. ^ Turner, R. L. (1931). "पञ्जा". A Comparative and Etymological Dictionary of the Nepali Language. London: K. Paul, Trench, Trübner. p. 359.
  4. ^ a b c d e f Glatston, A. R. (2021). "Introduction". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. xix–xxix. ISBN 978-0-12-823753-3.
  5. ^ Liddell, H. G. & Scott, R. (1940). "αἴλουρος". A Greek-English Lexicon (Revised and augmented ed.). Oxford: Clarendon Press.
  6. ^ Lewis, C. T. A. & Short, C. (1879). "fulgens". Latin Dictionary (Revised, enlarged, and in great part rewritten ed.). Oxford: Clarendon Press.
  7. ^ Cuvier, F. (1825). "Panda". In Geoffroy Saint-Hilaire, E.; Cuvier, F. (eds.). Histoire naturelle des mammifères, avec des figures originales, coloriées, dessinées d'après des animaux vivans: publié sous l'autorité de l'administration du Muséum d'Histoire naturelle. Vol. Tome 5. Paris: A. Belin. p. LII 1–3.
  8. ^ Cuvier, G. (1829). "Le Panda éclatant". Le règne animal distribué d'après son organisation. Vol. Tome 1. Chez Déterville, Paris. p. 138.
  9. ^ Hodgson, B. H. (1847). "On the Cat-toed subplantigrades". Journal of the Asiatic Society of Bengal. 16 (2): 1113–1128.
  10. ^ a b Groves, C. (2021). "The taxonomy and phylogeny of Ailurus". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 95–117. ISBN 978-0-12-823753-3.
  11. ^ Hu, Y.; Thapa, A.; Fan, H.; Ma, T.; Wu, Q.; Ma, S.; Zhang, D.; Wang, B.; Li, M.; Yan, L. & Wei, F. (2020). "Genomic evidence for two phylogenetic species and long-term population bottlenecks in red pandas". Science Advances. 6 (9): eaax5751. Bibcode:2020SciA....6.5751H. doi:10.1126/sciadv.aax5751. PMC 7043915. PMID 32133395.
  12. ^ Joshi, B. D.; Dalui, S.; Singh, S. K.; Mukherjee, T.; Chandra, K.; Sharma, L. K. & Thakur, M. (2021). "Siang river in Arunachal Pradesh splits Red Panda into two phylogenetic species". Mammalian Biology. 101 (1): 121–124. doi:10.1007/s42991-020-00094-y. S2CID 231811193.
  13. ^ Dalui, S.; Singh, S. K.; Joshi, B. D.; Ghosh, A.; Basu, S.; Khatri, H.; Sharma, L. K.; Chandra, K. & Thakur, M. (2021). "Geological and Pleistocene glaciations explain the demography and disjunct distribution of Red Panda (A. fulgens) in eastern Himalayas". Scientific Reports. 11 (1): 65. doi:10.1038/s41598-020-80586-6. PMC 7794540. PMID 33420314.
  14. ^ Pecon-Slattery, J. & O'Brien, S. J. (1995). "Molecular phylogeny of the red panda (Ailurus fulgens)". The Journal of Heredity. 86 (6): 413–422. doi:10.1093/oxfordjournals.jhered.a111615. PMID 8568209.
  15. ^ Flynn, J. J.; Finarelli, J. A.; Zehr, S.; Hsu, J. & Nedbal, M. A. (2005). "Molecular phylogeny of the Carnivora (Mammalia): Assessing the impact of increased sampling on resolving enigmatic relationships". Systematic Biology. 54 (2): 317–337. doi:10.1080/10635150590923326. PMID 16012099.
  16. ^ Law, C. J.; Slater, G. J. & Mehta, R. S. (2018). "Lineage Diversity and Size Disparity in Musteloidea: Testing Patterns of Adaptive Radiation Using Molecular and Fossil-Based Methods". Systematic Biology. 67 (1): 127–144. doi:10.1093/sysbio/syx047. PMID 28472434.
  17. ^ a b c Salesa, M. J.; Peigné, S.; Antón, M. & Morales, J. (2021). "The taxonomy and phylogeny of Ailurus". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 15–29. ISBN 978-0-12-823753-3.
  18. ^ a b Salesa, M. J.; Mauricio, A.; Peigné, S. & Morales, J. (2006). "Evidence of a false thumb in a fossil carnivore clarifies the evolution of pandas". PNAS. 103 (2): 379–382. Bibcode:2006PNAS..103..379S. doi:10.1073/pnas.0504899102. PMC 1326154. PMID 16387860.
  19. ^ a b c d e Wallace, S. C. & Lyon, L. (2021). "Systemic revision of the Ailurinae (Mammalia: Carnivora: Ailuridae): with a new species from North America". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 31–52. ISBN 978-0-12-823753-3.
  20. ^ a b Wallace, S. C. & Wang, X. (2004). "Two new carnivores from an unusual late Tertiary forest biota in eastern North America". Nature. 431 (7008): 556–559. Bibcode:2004Natur.431..556W. doi:10.1038/nature02819. PMID 15457257. S2CID 4432191.
  21. ^ Tedford, R. H. & Gustafson, E. P. (1977). "First North American record of the extinct panda Parailurus". Nature. 265 (5595): 621–623. Bibcode:1977Natur.265..621T. doi:10.1038/265621a0. S2CID 4214900.
  22. ^ Sotnikova, M. V. (2008). "A new species of lesser panda Parailurus (Mammalia, Carnivora) from the Pliocene of Transbaikalia (Russia) and some aspects of ailurine phylogeny". Paleontological Journal. 42 (1): 90–99. doi:10.1007/S11492-008-1015-X. S2CID 82000411.
  23. ^ Su, B.; Fu, Y.; Wang, Y.; Jin, L. & Chakraborty, R. (2001). "Genetic diversity and population history of the Red Panda (Ailurus fulgens) as inferred from mitochondrial DNA sequence variations". Molecular Biology and Evolution. 18 (6): 1070–1076. doi:10.1093/oxfordjournals.molbev.a003878. PMID 11371595.
  24. ^ Hu, Y.; Wu, Q.; Ma, S.; Ma, T.; Shan, L.; Wang, X.; Nie, Y.; Ning, Z.; Yan, L.; Xiu, Y. & Wei, F. (2017). "Comparative genomics reveals convergent evolution between the bamboo-eating giant and red pandas". Proceedings of the National Academy of Sciences. 114 (5): 1081–1086. doi:10.1073/pnas.1613870114. PMC 5293045. PMID 28096377.
  25. ^ Shan, L.; Wu, Q.; Wange, L.; Zhang, L. & Wei, F. (2017). "Lineage-specific evolution of bitter taste receptor genes in the giant and red pandas implies dietary adaptation". Integrative Zoology. 13 (2): 152–159. doi:10.1111/1749-4877.12291. PMC 5873442. PMID 29168616.
  26. ^ a b c d e f g Fisher, R. E. (2021). "Red Panda anatomy". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 81–93. doi:10.1016/B978-0-12-823753-3.00030-2. ISBN 978-0-12-823753-3. S2CID 243824295.
  27. ^ a b c d e f g h i j k Roberts, M. S. & Gittleman, J. L. (1984). "Ailurus fulgens" (PDF). Mammalian Species. 222 (222): 1–8. doi:10.2307/3503840. JSTOR 3503840.
  28. ^ Makungu, M.; du Plessis, W. M.; Groenewald, H. B.; Barrows, M. & Koeppel, K. N. (2015). "Morphology of the pelvis and hind limb of the Red Panda (Ailurus fulgens) evidenced by gross osteology, radiography and computed tomography". Anatomia, Histologia, Embryologia. 44 (6): 410–421. doi:10.1111/ahe.12152. hdl:2263/50447. PMID 25308447. S2CID 13035672.
  29. ^ Antón, M.; Salesa, M. J.; Pastor, J. F.; Peigné, S. & Morales, J. (2006). "Implications of the functional anatomy of the hand and forearm of Ailurus fulgens (Carnivora, Ailuridae) for the evolution of the 'false-thumb' in pandas". Journal of Anatomy. 209 (6): 757–764. doi:10.1111/j.1469-7580.2006.00649.x. PMC 2049003. PMID 17118063.
  30. ^ Kong, F.; Zhao, J.; Han, S.; Zeng, B.; Yang, J.; Si, X.; Yang, B.; Yang, M.; Xu, H. & Li, Y. (2014). "Characterization of the gut microbiota in the red panda (Ailurus fulgens)". PLOS ONE. 9 (2): e87885. Bibcode:2014PLoSO...987885K. doi:10.1371/journal.pone.0087885. PMC 3912123. PMID 24498390.
  31. ^ a b Kandel, K.; Huettmann, F.; Suwal, M. K.; Regmi, G. R.; Nijman, V.; Nekaris, K. A. I.; Lama, S. T.; Thapa, A.; Sharma, H. P. & Subedi, T. R. (2015). "Rapid multi-nation distribution assessment of a charismatic conservation species using open access ensemble model GIS predictions: Red Panda (Ailurus fulgens) in the Hindu-Kush Himalaya region". Biological Conservation. 181: 150–161. doi:10.1016/j.biocon.2014.10.007.
  32. ^ a b Mallick, J. K. (2010). "Status of Red Panda Ailurus fulgens in Neora Valley National Park, Darjeeling District, West Bengal, India". Small Carnivore Conservation. 43: 30–36.
  33. ^ a b c Dorji, S.; Rajaratnam, R. & Vernes, K. (2012). "The Vulnerable Red Panda Ailurus fulgens in Bhutan: distribution, conservation, status and management recommendations". Oryx. 46 (4): 536–543. doi:10.1017/S0030605311000780. S2CID 84332758.
  34. ^ a b c Thapa, K.; Thapa, G. J.; Bista, D.; Jnawali, S. R.; Acharya, K. P.; Khanal, K.; Kandel, R. C.; Karki Thapa, M.; Shrestha, S.; Lama, S. T. & Sapkota, N. S. (2020). "Landscape variables affecting the Himalayan Red Panda Ailurus fulgens occupancy in wet season along the mountains in Nepal". PLOS ONE. 15 (12): e0243450. Bibcode:2020PLoSO..1543450T. doi:10.1371/journal.pone.0243450. PMC 7740865. PMID 33306732.
  35. ^ Dong, X.; Zhang, J.; Gu, X.; Wang, Y.; Bai, W. & Huang, Q. (2021). "Evaluating habitat suitability and potential dispersal corridors across the distribution landscape of the Chinese Red Panda (Ailurus styani) in Sichuan, China". Global Ecology and Conservation. 28: e01705. doi:10.1016/j.gecco.2021.e01705.
  36. ^ a b c Lin, A. K.; Lwin, N.; Aung, S. S.; Oo, W. N.; Lum, L. Z. & Grindley, M. (2021). "The conservation status of Red Panda in north-east Myanmar". In Glatston, A. R. (ed.). Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 475–488. ISBN 9780128237540.
  37. ^ Shrestha, S.; Lama, S.; Sherpa, A. P.; Ghale, D. & Lama, S. T. (2021). "The endangered Himalayan Red Panda: first photographic evidence from its westernmost distributon range". Journal of Threatened Taxa. 13 (5): 18156–18163. doi:10.11609/jott.6100.13.5.18156-18163.
  38. ^ a b c Bista, D.; Baxter, G. S.; Hudson, N. J.; Lama, S. T.; Weerman, J. & Murray, P. J. (2021). "Movement and dispersal of a habitat specialist in human-dominated landscapes: a case study of the Red Panda". Movement Ecology. 9 (1): 62. doi:10.1186/s40462-021-00297-z. PMC 8670026. PMID 34906253.
  39. ^ a b Dalui, S.; Khatri, H.; Singh, S. K.; Basu, S.; Ghosh, A.; Mukherjee, T.; Sharma, L. K.; Singh, R.; Chandra, K. & Thakur, M. (2020). "Fine-scale landscape genetics unveiling contemporary asymmetric movement of Red Panda (Ailurus fulgens) in Kangchenjunga landscape, India". Scientific Reports. 10 (1): 15446. Bibcode:2020NatSR..1015446D. doi:10.1038/s41598-020-72427-3. PMC 7508845. PMID 32963325.
  40. ^ a b c Panthi, S.; Aryal, A.; Raubenheimer, D.; Lord, J. & Adhikari, B. (2012). "Summer diet and distribution of the Red Panda (Ailurus fulgens fulgens) in Dhorpatan Hunting Reserve, Nepal". Zoological Studies. 51 (5): 701–709.
  41. ^ Khatiwara, S. & Srivastava, T. (2014). "Red Panda Ailurus fulgens and other small carnivores in Kyongnosla Alpine Sanctuary, East Sikkim, India". Small Carnivore Conservation. 50: 35–38.
  42. ^ Bashir, T.; Bhattacharya, T.; Poudyal, K. & Sathyakumar, S. (2019). "First camera trap record of Red Panda Ailurus fulgens (Cuvier, 1825) (Mammalia: Carnivora: Ailuridae) from Khangchendzonga, Sikkim, India". Journal of Threatened Taxa. 11 (8): 14056–14061. doi:10.11609/jott.4626.11.8.14056-14061.
  43. ^ a b Chakraborty, R.; Nahmo, L. T.; Dutta, P. K.; Srivastava, T.; Mazumdar, K. & Dorji, D. (2015). "Status, abundance, and habitat associations of the Red Panda (Ailurus fulgens) in Pangchen Valley, Arunachal Pradesh, India". Mammalia. 79 (1): 25–32. doi:10.1515/mammalia-2013-0105. S2CID 87668179.
  44. ^ Megha, M.; Christi, S.; Kapoor, M.; Gopal, R. & Solanki, R. (2021). "Photographic evidence of Red Panda Ailurus fulgens Cuvier, 1825 from West Kameng and Shi-Yomi districts of Arunachal Pradesh, India". Journal of Threatened Taxa. 13 (9): 19254–19262. doi:10.11609/jott.6666.13.9.19254-19262.
  45. ^ Dong, X.; Zhang, J.; Gu, X.; Wang, Y.; Bai, W. & Huang, Q. (2021). "Evaluating habitat suitability and potential dispersal corridors across the distribution landscape of the Chinese red panda (Ailurus styani) in Sichuan, China". Global Ecology and Conservation. 28: e01705. doi:10.1016/j.gecco.2021.e01705.
  46. ^ a b c Wei, F.; Feng, Z.; Wang, Z. & Hu, J. (1999). "Current distribution, status and conservation of wild Red Pandas Ailurus fulgens in China". Biological Conservation. 89 (3): 285–291. doi:10.1016/S0006-3207(98)00156-6.
  47. ^ Li, F.; Huang, X. Y.; Zhang, X. C.; Zhao, X. X.; Yang, J. H. & Chan, B. P. L. (2019). "Mammals of Tengchong Section of Gaoligongshan National Nature Reserve in Yunnan Province, China". Journal of Threatened Taxa. 11 (11): 14402–14414. doi:10.11609/jott.4439.11.11.14402-14414.
  48. ^ a b c Wei, F.; Feng, Z.; Wang, Z. & Hu, J. (2000). "Habitat use and separation between the Giant Panda and the Red Panda". Journal of Mammalogy. 81 (2): 448–455. doi:10.1644/1545-1542(2000)081<0448:HUASBT>2.0.CO;2.
  49. ^ a b Dorji, S.; Vernes, K. & Rajaratnam, R. (2011). "Habitat correlates of the Red Panda in the temperate forests of Bhutan". PLOS ONE. 6 (10): e26483. Bibcode:2011PLoSO...626483D. doi:10.1371/journal.pone.0026483. PMC 3198399. PMID 22039497.
  50. ^ Dendup, P.; Lham, C.; Wangchuk, J. & Tshering, K. (2018). "Winter habitat preferences of Endangered Red Panda (Ailurus fulgens) in the Forest Research Preserve of Ugyen Wangchuck Institute for Conservation and Environmental Research, Bumthang, Bhutan". Journal of the Bhutan Ecological Society. 3: 1–13.
  51. ^ a b Dendup, P.; Humle, T.; Bista, D.; Penjor, U.; Lham, C. & Gyeltshen, J. (2020). "Habitat requirements of the Himalayan Red Panda (Ailurus fulgens) and threat analysis in Jigme Dorji National Park, Bhutan". Ecology and Evolution. 10 (17): 9444–9453. doi:10.1002/ece3.6632. PMC 7487235. PMID 32953073.
  52. ^ a b Zhang, Z.; Wei, F.; Li, M. & Hu, J. (2006). "Winter microhabitat separation between Giant and Red Pandas in Bashania faberi Bamboo forest in Fengtongzhai Nature Reserve". The Journal of Wildlife Management. 70 (1): 231–235. doi:10.2193/0022-541X(2006)70[231:WMSBGA]2.0.CO;2.
  53. ^ Dendup, P.; Lham, C.; Wangchuk, J. & Tshering, K. (2018). "Winter habitat preferences of Endangered Red Panda (Ailurus fulgens) in the Forest Research Preserve of Ugyen Wangchuck Institute for Conservation and Environmental Research, Bumthang, Bhutan". Journal of the Bhutan Ecological Society. 3: 1–13.
  54. ^ Zhou, X.; Jiao, H.; Dou, Y.; Aryal, A.; Hu, J.; Hu, J. & Meng, X. (2013). "The winter habitat selection of Red Panda (Ailurus fulgens) in the Meigu Dafengding National Nature Reserve, China". Current Science. 105 (10): 1425–1429.
  55. ^ Liu, X.; Teng, L.; Ding, Y. & Liu, Z. (2021). "Habitat selection by Red Panda (Ailurus fulgens fulgens) in Gaoligongshan Nature Reserve, China" (PDF). Pakistan Journal of Zoology. doi:10.17582/journal.pjz/20190726090725. S2CID 238963119.
  56. ^ a b Gebauer, A. (2021). "The early days: maternal behaviour and infant development". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 149–179. ISBN 978-0-12-823753-3.
  57. ^ Karki, S.; Maraseni, T.; Mackey, B.; Bista, D.; Lama, S. T.; Gautam, A. P.; Sherpa, A. P.; Koju, A. P.; Koju, U.; Shrestha, A. & Cadman, T. (2021). "Reaching over the gap: A review of trends in and status of red panda research over 193 years (1827–2020)". Science of the Total Environment. 781: 146659. Bibcode:2021ScTEn.781n6659K. doi:10.1016/j.scitotenv.2021.146659. PMID 33794452. S2CID 232763016.
  58. ^ a b Reid, D. G.; Jinchu, H. & Yan, H. (1991). "Ecology of the Red Panda Ailurus fulgens in the Wolong Reserve, China". Journal of Zoology. 225 (3): 347–364. doi:10.1111/j.1469-7998.1991.tb03821.x.
  59. ^ a b Bista, D.; Baxter, G. S.; Hudson, N. J.; Lama, S. T. & Murray, P. J. (2021). "Effect of disturbances and habitat fragmentation on an arboreal habitat specialist mammal using GPS telemetry: a case of the red panda". Landscape Ecology: 1–15. doi:10.1007/s10980-021-01357-w. PMC 8542365. PMID 34720409.
  60. ^ Nijboer, J. & Dierenfeld, E. S. (2021). "Red panda nutrition: how to feed a vegetarian carnivore". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 225–238. ISBN 978-0-12-823753-3.
  61. ^ Wei, F.; Thapa, A.; Hu, Y. & Zhang, Z. (2021). "Red Panda ecology". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 329–351. ISBN 978-0-12-823753-3.
  62. ^ WFO (2022). "Arundinaria maling Gamble". World Flora Online. Retrieved 8 February 2022.
  63. ^ WFO (2022). "Arundinaria aristata Gamble". World Flora Online. Retrieved 8 February 2022.
  64. ^ Pradhan, S.; Saha, G. K. & Khan, J. A. (2001). "Ecology of the Red Panda Ailurus fulgens in the Singhalila National Park, Darjeeling, India". Biological Conservation. 98: 11–18. doi:10.1016/S0006-3207(00)00079-3.
  65. ^ Roka, B.; Jha, A. K. & Chhetri, D. R. (2021). "A study on plant preferences of Red Panda (Ailurus fulgens) in the wild habitat: foundation for the conservation of the species". Acta Biologica Sibirica. 7: 425–439. doi:10.3897/abs.7.e71816. S2CID 244942192.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  66. ^ Sharma, H. R.; Swenson, J. E. & Belant, J. (2014). "Seasonal food habits of the Red Panda (Ailurus fulgens) in Rara National Park, Nepal". Hystrix. 25 (1): 47–50. doi:10.4404/hystrix-25.1-9033.
  67. ^ Wei, F.; Feng, Z.; Wang, Z.; Zhou, A. & Hu, J. (1999). "Use of the nutrients in bamboo by the Red Panda Ailurus fulgens". Journal of Zoology. 248 (4): 535–541. doi:10.1111/j.1469-7998.1999.tb01053.x.
  68. ^ Qi, D.; Zhou, H.; Wei, W.; Lei, M.; Yuan, S.; Qi, D. & Zhang, Z. (2016). "Vocal repertoire of adult captive red pandas (Ailurus fulgens)". Animal Biology. 66 (2): 145–155. doi:10.1163/15707563-00002493.
  69. ^ a b c Curry, E. (2021). "Reproductive biology of the Red Panda". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 119–138. ISBN 978-0-12-823753-3.
  70. ^ Roberts, M. S. & Kessler, D. S. (1979). "Reproduction in Red pandas, Ailurus fulgens (Carnivora : Ailuropodidae)". Journal of Zoology. 188 (2): 235–249. doi:10.1111/j.1469-7998.1979.tb03402.x.
  71. ^ Lama, S. T.; Lama, R. P.; Regmi, G. R. & Ghimire, T. R. (2015). "Prevalence of intestinal parasitic infections in free-ranging Red Panda Ailurus fulgens Cuvier, 1825 (Mammalia: Carnivora: Ailuridae) in Nepal". Journal of Threatened Taxa. 7 (8): 7460–7464. doi:10.11609/JoTT.o4208.7460-4.
  72. ^ Bista, D.; Shrestha, S.; Kunwar, A. J.; Acharya, S.; Jnawali, S. R. & Acharya, K. P. (2017). "Status of gastrointestinal parasites in Red Panda of Nepal". PeerJ. 5: e3767. doi:10.7717/peerj.3767. PMC 5591639. PMID 28894643.
  73. ^ Sharma, H. P. & Achhami, B. (2021). "Gastro‐intestinal parasites of sympatric Red Panda and livestock in protected areas of Nepal". Veterinary Medicine and Science. doi:10.1002/vms3.651. PMID 34599791. S2CID 238250774.
  74. ^ Kearns, K. S.; Pollock, C. G. & Ramsay, E. C. (1999). "Dermatophytosis in Red Pandas (Ailurus fulgens fulgens): a review of 14 cases". Journal of Zoo and Wildlife Medicine. 30 (4): 561–563. JSTOR 20095922. PMID 10749446.
  75. ^ Huckins, G. L.; Eshar, D.; Schwartz, D.; Morton, M.; Herrin, B. H.; Cerezo, A.; Yabsley, M. J. & Schneider, S. M. (2019). "Trypanosoma cruzi infection in a zoo-housed Red Panda in Kansas". Journal of Veterinary Diagnostic Investigation. 31 (5): 752–755. doi:10.1177/1040638719865926. PMC 6727118. PMID 31342874.
  76. ^ Alex, C. E.; Kubiski, S. V.; Li, L.; Sadeghi, M.; Wack, R. F.; McCarthy, M. A.; Pesavento, J. B.; Delwart, E. & Pesavento, P. A. (2018). "Amdoparvovirus infection in Red Pandas (Ailurus fulgens)". Veterinary Pathology. 55 (4): 552–561. doi:10.1177/0300985818758470. PMID 29433401.
  77. ^ Yang, Y.; Dong, H.; Su, R.; Li, T.; Jiang, N.; Su, C. & Zhang, L. (2019). "Evidence of Red Panda as an intermediate host of Toxoplasma gondii and Sarcocystis species". International Journal for Parasitology: Parasites and Wildlife. 8: 188–191. doi:10.1016/j.ijppaw.2019.02.006. PMC 6403407. PMID 30891398.
  78. ^ Liu, S.; Li, Y.; Yue, C.; Zhang, D.; Su, X.; Yan, X.; Yang, K.; Chen, X.; Zhuo, G.; Cai, T.; Liu, J.; Peng, X. & Huo, R. (2020). "Isolation and characterization of uropathogenic Escherichia coli (UPEC) from Red Panda (Ailurus fulgens)". BMC Veterinary Research. 16 (1): 404. doi:10.1186/s12917-020-02624-9. PMC 7590469. PMID 33109179.
  79. ^ a b Glatson, A. R. (1994). "The Red Panda or Lesser Panda (Ailurus fulgens)" (PDF). Status Survey and Conservation Action Plan for Procyonids and Ailurids. The Red Panda, Olingos, Coatis, Raccoons, and their Relatives. Gland, Switzerland: IUCN/SSC Mustelid, Viverrid, and Procyonid Specialist Group. pp. viii, 8–11, 19–21. ISBN 2-8317-0046-9.
  80. ^ Yonzon, P. B. & Hunter, M. L. Jr. (1991). "Conservation of the Red Panda Ailurus fulgens". Biological Conservation. 57 (1): 1–11. doi:10.1016/0006-3207(91)90046-C.
  81. ^ Acharya, K. P.; Shrestha, S.; Paudel, P. K.; Sherpa, A. P.; Jnawali, S. R.; Acharya, S. & Bista, D. (2018). "Pervasive human disturbance on habitats of endangered Red Panda Ailurus fulgens in the central Himalaya". Global Ecology and Conservation. 15: e00420. doi:10.1016/j.gecco.2018.e00420. S2CID 92988737.
  82. ^ Ghimire, G.; Pearch, M.; Baral, B.; Thapa, B. & Baral, R. (2019). "The first photographic record of the Red Panda Ailurus fulgens (Cuvier, 1825) from Lamjung District outside Annapurna Conservation Area, Nepal". Journal of Threatened Taxa. 11 (12): 14576–14581. doi:10.11609/jott.4828.11.12.14576-14581.
  83. ^ Bista, D.; Baxter, G. S. & Murray, P. J. (2020). "What is driving the increased demand for red panda pelts?". Human Dimensions of Wildlife. 25 (4): 324–338. doi:10.1080/10871209.2020.1728788. S2CID 213958948.
  84. ^ Xu, L. & Guan, J. (2018). Red Panda market research findings in China (PDF). Cambridge: Traffic.
  85. ^ Datta, A.; Naniwadekar, R. & Anand, M. O. (2008). "Occurrence and conservation status of small carnivores in two protected areas in Arunachal Pradesh, north-east India". Small Carnivore Conservation. 39: 1–10.
  86. ^ Lwin, Y. H.; Wang, L.; Li, G.; Maung, K. W.; Swa, K. & Quan, R. C. (2021). "Diversity, distribution and conservation of large mammals in northern Myanmar". Global Ecology and Conservation. 29: e01736. doi:10.1016/j.gecco.2021.e01736.
  87. ^ Li, X.; Bleisch, W. V.; Liu, X. & Jiang, X. (2021). "Camera-trap surveys reveal high diversity of mammals and pheasants in Medog, Tibet". Oryx. 55 (2): 177–180. doi:10.1017/S0030605319001467.
  88. ^ a b Wei, F.; Zhang, Z.; Thapa, A.; Zhijin, L. & Hu, Y. (2021). "Conservation initiatives in China". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 509–520. doi:10.1016/B978-0-12-823753-3.00021-1. ISBN 978-0-12-823753-3. S2CID 243813871.
  89. ^ Thapa, A.; Hu, Y. & Wei, F. (2018). "The endangered Red Panda (Ailurus fulgens): Ecology and conservation approaches across the entire range". Biological Conservation. 220: 112–121. doi:10.1016/j.biocon.2018.02.014.
  90. ^ Bista, D. (2018). "Communities in frontline in Red Panda conservation, eastern Nepal" (PDF). The Himalayan Naturalist. 1 (1): 11–12.
  91. ^ Sherpa, A. P.; Lama, S. T.; Shrestah, S.; Williams, B. & Bista, D. (2021). "Red Pandas in Nepal: community-based approach to landscape-level conservation". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 495–508. doi:10.1016/B978-0-12-823753-3.00019-3. ISBN 978-0-12-823753-3. S2CID 243829246.
  92. ^ Millar, J. & Tenzing, K. (2021). "Transforming degraded rangelands and pastoralists' livelihoods in eastern Bhutan". Mountain Research and Development. 41 (4): D1–D7. doi:10.1659/MRD-JOURNAL-D-21-00025.1.
  93. ^ Jones, M. L. (2021). "A brief history of the Red Panda in captivity". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 181–199. doi:10.1016/B978-0-12-823753-3.00026-0. ISBN 978-0-12-823753-3. S2CID 243805749.
  94. ^ Lewis, M. (2011). "Birth and mother rearing of Nepalese red pandas Ailurus fulgens fulgens at the Taronga Conservation Society Australia". International Zoo Yearbook. 45 (1): 250–258. doi:10.1111/j.1748-1090.2011.00135.x.
  95. ^ Kappelhof, J. & Weerman, J. (2020). "The development of the Red panda Ailurus fulgens EEP: from a failing captive population to a stable population that provides effective support to in situ conservation". International Zoo Yearbook. 54 (1): 102–112. doi:10.1111/izy.12278.
  96. ^ Tanaka, A. & Ogura, T. (2018). "Current husbandry situation of Red Pandas in Japan". Zoo Biology. 37 (2): 107–114. doi:10.1002/zoo.21407. PMID 29512188.
  97. ^ Kumar, A.; Rai, U.; Roka, B.; Jha, A. K. & Reddy, P. A. (2016). "Genetic assessment of captive red panda (Ailurus fulgens) population". SpringerPlus. 5 (1): 1750. doi:10.1186/s40064-016-3437-1. PMC 5055525. PMID 27795893.
  98. ^ a b Glatston, A. R. & Gebauer, A. (2021). "People and Red Pandas: the Red Panda's role in economy and culture". In Glatston, A. R. (ed.). Red Panda: Biology and Conservation of the First Panda (Second ed.). London: Academic Press. pp. 1–14. doi:10.1016/B978-0-12-823753-3.00002-8. ISBN 9780128237540. S2CID 243805192.
  99. ^ Lowther, D. A. (2021). "The first painting of the Red Panda (Ailurus fulgens) in Europe? Natural history and artistic patronage in early nineteenth-century India". Archives of Natural History. 48 (2): 368–376. doi:10.3366/anh.2021.0728. S2CID 244938631.

External links